19
materials Article Synthesis and Characterization of Bio-Based Polyesters: Poly(2-methyl-1,3-propylene-2,5-furanoate), Poly(isosorbide-2,5-furanoate), Poly(1,4-cyclohexanedimethylene-2,5-furanoate) Zoi Terzopoulou 1 , Nejib Kasmi 1 , Vasilios Tsanaktsis 1 , Nikolaos Doulakas 1 , Dimitrios N. Bikiaris 1, *, Dimitris S. Achilias 1 ID and George Z. Papageorgiou 2, * ID 1 Laboratory of Polymer Chemistry and Technology, Department of Chemistry, Aristotle University of Thessaloniki, Thessaloniki GR-541 24, Greece; [email protected] (Z.T.); [email protected] (N.K.); [email protected] (V.T.); [email protected] (N.D.); [email protected] (D.S.A.) 2 Department of Chemistry, University of Ioannina, Ioannina GR-54110, Greece * Correspondence: [email protected] (D.N.B.); [email protected] (G.Z.P.); Tel.: +30-231-0997812 (D.N.B.); +30-265-1008354 (G.Z.P.) Received: 15 June 2017; Accepted: 12 July 2017; Published: 14 July 2017 Abstract: In the present study, three new biobased furanoate polyesters with potential use in food packaging applications, named poly(isosorbide furanoate) (PIsF), poly(methyl-propylene furanoate) (PMePF) and poly(1,4-cyclohexane-dimethylene 2,5-furanoate) (PCHDMF) were synthesized. As monomers for the preparation of the polyesters, 2,5-furandicarboxylic acid (FDCA) and diols with irregular or complicated structure were used, including isosorbide (IS), 2-methyl-1,3-propanediol (MPD) and 1,4-cyclohexane-dimethanol (CHDM). The polymerization process was carried out via melt polycondensation method. The structural characteristics and thermal behavior of the polymers were studied. The kinetic fragility of the amorphous phase of the polymers was evaluated. The thermal degradation was studied by means of thermogravimetry and a pyrolysis Py-GC/MS (Pyrolysis-Gas Chromatography/Mass Spectroscopy) system to estimate the degradation mechanism. Keywords: furanoate polyesters; complicated structure diols; synthesis; thermal properties 1. Introduction The exploitation of fossil based resources and mainly of petroleum to produce petrochemical polymers has as a result the shrinkage of the above. Furfural (F) and 5-hydroxymethylfurfural (HMF) can be isolated from carbohydrates, and they constitute precursor compounds of renewable monomers such as 2,5-diformylfuran (DFF) [1], 5-hydroxymethyl-2-furancarboxylic acid (HMFCA) [2] and FDCA [3]. Poly(ethylene furanoate) (PEF) gains ground, against poly(ethylene terephthalate) (PET), especially in food packaging applications, due to renewable sourcing of the monomers and the largely improved barrier, mechanical, and thermal properties compared to PET [4]. PET is one of the most widely used thermoplastics in the world, especially in bottle and fiber production, and its popularity derives from its versatility, since it can be produced in different molecular weights and easily manipulated crystallinity [5]. Nowadays, there is an increased interest to produce new furanoate polyesters using alternative diols to ethylene glycol [68]. Besides PEF, poly(propylene furanoate) (PPF) as well as other furanoate polyesters are expected to get their share in the market of food packaging and fibers [7]. Both industry and researchers are showing increased interest in the utilization of furan-based monomers for the replacement of alipharomatic polyesters widely used Materials 2017, 10, 801; doi:10.3390/ma10070801 www.mdpi.com/journal/materials

Poly(2-methyl-1,3-propylene-2,5-furanoate), Poly ... · irregular or complicated structure were used, including isosorbide (IS), 2-methyl-1,3-propanediol (MPD) and 1,4-cyclohexane-dimethanol

Embed Size (px)

Citation preview

materials

Article

Synthesis and Characterization of Bio-Based Polyesters:Poly(2-methyl-1,3-propylene-2,5-furanoate),Poly(isosorbide-2,5-furanoate),Poly(1,4-cyclohexanedimethylene-2,5-furanoate)

Zoi Terzopoulou 1, Nejib Kasmi 1, Vasilios Tsanaktsis 1, Nikolaos Doulakas 1,Dimitrios N. Bikiaris 1,*, Dimitris S. Achilias 1 ID and George Z. Papageorgiou 2,* ID

1 Laboratory of Polymer Chemistry and Technology, Department of Chemistry,Aristotle University of Thessaloniki, Thessaloniki GR-541 24, Greece; [email protected] (Z.T.);[email protected] (N.K.); [email protected] (V.T.); [email protected] (N.D.);[email protected] (D.S.A.)

2 Department of Chemistry, University of Ioannina, Ioannina GR-54110, Greece* Correspondence: [email protected] (D.N.B.); [email protected] (G.Z.P.);

Tel.: +30-231-0997812 (D.N.B.); +30-265-1008354 (G.Z.P.)

Received: 15 June 2017; Accepted: 12 July 2017; Published: 14 July 2017

Abstract: In the present study, three new biobased furanoate polyesters with potential use in foodpackaging applications, named poly(isosorbide furanoate) (PIsF), poly(methyl-propylene furanoate)(PMePF) and poly(1,4-cyclohexane-dimethylene 2,5-furanoate) (PCHDMF) were synthesized.As monomers for the preparation of the polyesters, 2,5-furandicarboxylic acid (FDCA) and diols withirregular or complicated structure were used, including isosorbide (IS), 2-methyl-1,3-propanediol(MPD) and 1,4-cyclohexane-dimethanol (CHDM). The polymerization process was carried outvia melt polycondensation method. The structural characteristics and thermal behavior of thepolymers were studied. The kinetic fragility of the amorphous phase of the polymers was evaluated.The thermal degradation was studied by means of thermogravimetry and a pyrolysis Py-GC/MS(Pyrolysis-Gas Chromatography/Mass Spectroscopy) system to estimate the degradation mechanism.

Keywords: furanoate polyesters; complicated structure diols; synthesis; thermal properties

1. Introduction

The exploitation of fossil based resources and mainly of petroleum to produce petrochemicalpolymers has as a result the shrinkage of the above. Furfural (F) and 5-hydroxymethylfurfural(HMF) can be isolated from carbohydrates, and they constitute precursor compounds of renewablemonomers such as 2,5-diformylfuran (DFF) [1], 5-hydroxymethyl-2-furancarboxylic acid (HMFCA) [2]and FDCA [3]. Poly(ethylene furanoate) (PEF) gains ground, against poly(ethylene terephthalate)(PET), especially in food packaging applications, due to renewable sourcing of the monomers andthe largely improved barrier, mechanical, and thermal properties compared to PET [4]. PET is oneof the most widely used thermoplastics in the world, especially in bottle and fiber production, andits popularity derives from its versatility, since it can be produced in different molecular weightsand easily manipulated crystallinity [5]. Nowadays, there is an increased interest to produce newfuranoate polyesters using alternative diols to ethylene glycol [6–8]. Besides PEF, poly(propylenefuranoate) (PPF) as well as other furanoate polyesters are expected to get their share in the marketof food packaging and fibers [7]. Both industry and researchers are showing increased interest in theutilization of furan-based monomers for the replacement of alipharomatic polyesters widely used

Materials 2017, 10, 801; doi:10.3390/ma10070801 www.mdpi.com/journal/materials

Materials 2017, 10, 801 2 of 19

for decades, such as PET, poly(trimethylene terephthalate) (PTT) and poly(butylene terephthalate)(PBT) [9].

IS constitutes a heterocyclic diol derived from glucose. Due to its complicated structure, rigidityand chirality, it forms amorphous polymers and co-polymers with terephthalate and succinate acidwith interesting thermal properties, high Tg values and/or special optical properties [10,11]. PIsF isconsidered an alternative to polycarbonate (PC), since in contrast to bisphenol A, isosorbide is anon-toxic compound [7]. PC is a widely used thermoplastic with excellent mechanical and thermalproperties with applications in automotive industry, data storage, electronics and baby products.By estimation, the production of bisphenol A in 2015 exceeded 5.4 million tons [5].

CHDM is produced by catalytic hydrogenation of dimethyl terephthalate (DMT). The reactionis conducted in two steps beginning with the conversion of DMT to the diester dimethyl1,4-cyclohexanedicarboxylate (DMCD). In the second step, DMCD is further hydrogenated to CHDM.The cis/trans ratio of CHDM is different depending on the catalyst [12,13]. CHDM is polymerized withterephthalic acid to form a linear polyester, poly(1,4-cyclohexylenedimethylene terephthalate), withimproved hydrolytic, chemical, electrical properties, as well as enhanced thermal properties [14]. As acomonomer, CHDM could enhance the mechanical properties, especially toughness, and crystallizationof PEF [15].

2-methyl-1,3-propanediol (MPD) is synthesized by isomerization of propylene oxide to allylalcohol, followed by hydrogenation [16], and is used as an intermediate for organic synthesis, as asolvent and in colorant products.

In the present study, three new furanoate polyesters were synthesized from the reaction ofdimethylester of furanoate dicarboxyl acid (DMFD) with MPD, CHDM and IS. For the synthesis ofPDMePF and PIsF, a typical two-step procedure was used. On the contrary, an innovate technique wasused to achieve the synthesis of PCHDMF due to the higher boiling point of CHDM, as in a previousstudy of ours [17]. The intrinsic viscosity of the polyesters was measured and the structure confirmationwas achieved by Nuclear Magnetic Resonance (NMR) spectroscopy. Thermal properties and thermalstability were determined by Differential Scanning Calorimetry (DSC) and Thermogravimetric Analysis(TGA), respectively, while the existence of degree of crystallinity was investigated by Wide Angle X-raydiffraction patterns (WAXD). Furthermore, the mechanism of thermal degradation during pyrolysis ofthe three polyesters was studied by Py/GC-MS.

2. Experimental

2.1. Materials

FDCA (purum 97%), MPD (99%, b.p. = 125 ◦C), CHDM (mixture of cis-trans, 99%, m.p. = 58 ◦Cand b.p. = 287 ◦C), Isosorbide (99%, m.p. = 63 ◦C and b.p. = 160 ◦C) and Tetrabutyltitanate (TBT)catalyst of analytical grade were purchased from Aldrich Co, Hamburg, Germany.

2.2. Polyester Synthesis

PIsF and PMePF were synthesized by the melt polycondensation process, including thetransesterification of dimethyl ester of FDCA (DMFD), and its reaction with isosorbide and2-methyl-1,3-propanediol, respectively, as described in our previous paper [17,18]. Briefly, themonomers were charged into the reaction apparatus with a molar ratio of diester/diol = 1/2 with400 ppm TBT. The reaction mixture was heated at 150 ◦C under argon atmosphere for 2 h, at 160 ◦Cfor additional 2 h, and finally at 170 ◦C for 1 h. This first step (transesterification) is considered to becomplete after the collection of almost all of the theoretical amount of CH3OH, which was removedfrom the reaction mixture by distillation and collected in a graduated cylinder. After this stage, thecorresponding bishydroxyalkylene-2,5-furandicarboxylate monomers have been formed. In the secondstage, these monomers reacted with DMFD in a molar ratio 1/1.05 at 150 ◦C under argon atmospherefor 2 h, at 160 ◦C for additional 2 h, and finally at 170 ◦C for 1 h. During this stage, methanol was

Materials 2017, 10, 801 3 of 19

also removed as byproduct. After that time, in the third step of polycondensation, a vacuum (5.0 Pa)was applied slowly over a period of about 30 min. The temperature was increased to 210 ◦C, and thepolymerization continued for 1 h at this temperature, at 220 ◦C for 1 h, and at 230 ◦C for 0.5 h using astirring speed of 720 rpm. For synthesis of PCHDMF these temperatures were 240, 250 and 260 ◦C,respectively. After the polycondensation reaction was completed, the polyesters were easily removed,milled, and washed with methanol.

In the case of PCHDMF, a variation of the two-stage melt polycondensation method was applied,due to the high boiling point of 1,4-CHDM, as described in our previous work [17,18].

2.3. Polyester Characterization

Intrinsic viscosity [IV] measurements of the prepared polyesters were performed using anUbbelohde viscometer at 30 ◦C in a mixture of phenol/1,1,2,2-tetrachloroethane (60/40 w/w).

1H-NMR spectra of polyesters were obtained with a Bruker spectrometer operating at a frequencyof 400 MHz for protons using deuteratedtrifluoroacetic acid (d-TFA) as solvent in order to preparesolutions of 5% w/v. The number of scans was 10 and the sweep width was 6 kHz.

Fourier transform infrared spectroscopy (FTIR) spectra of all the samples were obtained using aPerkin-Elmer FTIR spectrometer, model Spectrum One. The materials were in the form of thin filmswith thickness of approximately 15 mm. The IR spectra of these films were obtained in absorbancemode and in the spectral region of 400–4000 cm−1 using a resolution of 4 cm−1 and 64 co-added scans.

WAXD of the samples were recorded using a MiniFlex II XRD system from Rigaku Co., Tokyo,Japan, with CuKα radiation (λ = 0.154 nm) in the angle 2θ range from 5 to 60 degrees, afterpolyester annealing.

For thermal analysis measurements a Perkin–Elmer, Pyris Diamond DSC, coupled with anIntracooler 2P cooling accessory, was used. Samples of 10 ± 0.1 mg sealed in aluminum pans wereused, to test the thermal behavior of the polyesters. The samples were heated from −50 ◦C to 160 ◦C ina 20 mL/min flow of N2 with heating rate 20 ◦C/min in order to observe the melting temperature of theas received polyesters. The samples first were held at that temperature for 2 min, then quenched andrescanned again till 160 ◦C. TGA measurements were carried out by a STA 449C (Netzch-Gerätebau,GmbH, Selb, Germany) thermal analyzer from room temperature up to 600 ◦C with 20 ◦C/min heatingrate and 30 mL/min flow of N2 (99.9%).

Thermogravimetric analysis was carried out with a Setsys 16/18 TG-DTA (SetaramInstrumentation, Caluire-et-Cuire, France). Samples (4 ± 0.2 mg) were placed in alumina cruciblesand heated from ambient temperature to 650 ◦C at 20 ◦C/min in a 50 mL/min flow of N2; an emptyalumina crucible was used as reference.

For Py-GC/MS analysis of polyesters a very small amount of each material is “dropped” initiallyinto the “Double-Shot” EGA/PY-3030D Pyrolyzer (Frontier Laboratories Ltd., Fukushima, Japan)using a CGS-1050Ex (Frontier Laboratories Ltd, Fukushima, Japan) carrier gas selector. For EvolvedGas Analysis (EGA), the furnace temperature was programmed from 50 to 700 ◦C with a heatingrate of 20 ◦C/min, using He as purge gas and air as cooling gas. For pyrolysis analysis (flashpyrolysis) each sample was placed into the sample cup which afterwards fell free into the Pyrolyzerfurnace. The pre-selected pyrolysis temperature was 600 ◦C and the GC oven temperature was heatedfrom 70 to 300 ◦C at 10 ◦C/min. This temperature was selected based on the EGA. Sample vaporsgenerated in the furnace were split (at a ratio of 1/50), a portion moved to the column at a flow rate of1 mL/min, pressure 53.6 kPa and the remaining portion exited the system via the vent. The pyrolyzateswere separated using temperature programmed capillary column of a Shimadzu QP-2010 Ultra Plus(Shimadzu, Kioto, Japan) gas chromatograph and analyzed by the mass spectrometer MS-QP2010SE ofShimadzu (Shimadzu, Kioto, Japan) use 70 eV. Ultra ALLOY® metal capillary column from FrontierLaboratories LTD (Fukushima Japan) was used containing 5% diphenyl and 95% dimethylpolysiloxanestationary phase, column length 30 m and column ID 0.25 mm. For the mass spectrometer the followingconditions were used: Ion source heater 200 ◦C, interface temperature 320 ◦C, vacuum 10−4–100 Pa,

Materials 2017, 10, 801 4 of 19

m/z range 45–500 amu and scan speed 10,000. The chromatograph and spectra retrieved by eachexperiment are subject to further interpretation through Shimadzu and Frontier post-run software.

3. Results and Discussion

3.1. Polyester Synthesis

PIsF and PMePF were prepared by applying the two-stage melt polycondensation method(esterification and polycondensation) in a glass batch reactor [18]. The measured [η] values arepresented in Table 1. The PIsF sample showed [η] = 0.39 dL/g, while PMePF showed [η] = 0.42 dL/g.On the contrary, to synthesize PCHDMF, a variation of the typical method was applied, as wasdeveloped in our previous work [17,18], because 1,4-CHDM is hardly distillable, even after theapplication of vacuum. Using the traditional method, in the polycondensation stage, temperatureshigher than 250–270 ◦C should be applied prior, to remove the diol byproduct and to increase themolecular weight. However, at such high temperatures, the polyester decomposition would be veryextensive, leading to low molecular weight polyesters and coloration, which is also one of the mostserious problems in such polyesters. Applying the adjusted procedure, [η] = 0.52 dL/g for PCHDMFwas achieved. The measured intrinsic viscosities are typical for alipharomatic polyesters and this isproof that the followed procedure can be applied successfully for synthesis of such polyesters.

Table 1. Intrinsic viscosity values of the polyesters.

Polyester [η] (dL/g)

PMePF 0.42PIsF 0.39

PCHDMF 0.52

3.2. Structural Characterization

The structure of the prepared polyesters was verified with 1H-NMR spectroscopy. The spectraof PIsF, PMePF and PCHDMF are shown in Figure 1. At first sight, it is easy to attribute the “a”,“e” and “j” picks to the ring protons at 7.39, 7.46 and 7.44 ppm respectively (2 H, s), due to theirhighest π-deprotection. The PMePF spectra presents a double pick for “b” protons, which are the mostdeprotected aliphatic protons due to their location next to the ester bond (4 H, d). The “c” protoninteracts with 7 protons and gives a multiple peak at 2.58 ppm (1 H, m). The lowest deprotected arethe “d” methyl protons and appear at the value of 1.2 ppm (3 H, d).

Materials 2017, 10, 801 4 of 18

3. Results and Discussion

3.1. Polyester Synthesis

PIsF and PMePF were prepared by applying the two-stage melt polycondensation method (esterification and polycondensation) in a glass batch reactor [18]. The measured [η] values are presented in Table 1. The PIsF sample showed [η] = 0.39 dL/g, while PMePF showed [η] = 0.42 dL/g. On the contrary, to synthesize PCHDMF, a variation of the typical method was applied, as was developed in our previous work [17,18], because 1,4-CHDM is hardly distillable, even after the application of vacuum. Using the traditional method, in the polycondensation stage, temperatures higher than 250–270 °C should be applied prior, to remove the diol byproduct and to increase the molecular weight. However, at such high temperatures, the polyester decomposition would be very extensive, leading to low molecular weight polyesters and coloration, which is also one of the most serious problems in such polyesters. Applying the adjusted procedure, [η] = 0.52 dL/g for PCHDMF was achieved. The measured intrinsic viscosities are typical for alipharomatic polyesters and this is proof that the followed procedure can be applied successfully for synthesis of such polyesters.

Table 1. Intrinsic viscosity values of the polyesters.

Polyester [η] (dL/g)PMePF 0.42

PIsF 0.39PCHDMF 0.52

3.2. Structural Characterization

The structure of the prepared polyesters was verified with 1H-NMR spectroscopy. The spectra of PIsF, PMePF and PCHDMF are shown in Figure 1. At first sight, it is easy to attribute the “a”, “e” and “j” picks to the ring protons at 7.39, 7.46 and 7.44 ppm respectively (2 H, s), due to their highest π-deprotection. The PMePF spectra presents a double pick for “b” protons, which are the most deprotected aliphatic protons due to their location next to the ester bond (4 H, d). The “c” proton interacts with 7 protons and gives a multiple peak at 2.58 ppm (1 H, m). The lowest deprotected are the “d” methyl protons and appear at the value of 1.2 ppm (3 H, d).

(a)

Figure 1. Cont.

Materials 2017, 10, 801 5 of 19

Materials 2017, 10, 801 5 of 18

14 12 10 8 6 4 2 0 -2

ppm

TFA

a b

c

d

e f g hi

j

kl m

PMePF

PIsF

PCHDMF

(b)

Figure 1. (a) Structures; and (b) 1H-NMR spectra of PMePF, PIsF and PCHDMF polyesters.

The isosorbide part of PIsF shows a more complicated spectrum due to its spatial configuration. The most deprotected are the “f” protons owing to their “next to ester bond” location and appear at 5.76 ppm (2 H, q). The endo and exo “g” protons are more deprotected than the exo and endo “h” protons, due to the presence of endo and exo ester bond and appear at 5.76 ppm (2 H, t) and 5.45 ppm (2 H, t), respectively. The “i” protons show a different peak, 4.33 ppm (1 H, d) and 4.50 ppm (1 H, d), as they are characterized as exo protons and are in varying degree deprotected by the endo and exo ester bonds, respectively.

At the PCHDMF spectrum, the “k” protons correspond to two double peaks at 4.38 (2 H, d) and 4.5 ppm (2 H, d), respectively. This phenomenon is attributed to the fact that the used CHMD was a cis-trans mixture. Thus, the axial “k” protons at trans case interact different than the axial-equatorial “k” protons at cis case. The “l” protons show a multiple peak at 2.00 ppm (2H, m) and the “m” protons show a double peak at 1.20 ppm (2 H, d). The absent of any other peaks indicates that polyester with high purity have been prepared.

FTIR spectra of the three synthesized polyesters are presented in Figure 2. The spectrum of PMePF shows characteristic peaks at 3550 cm−1 due to the stretching of the O–H stretch of the carboxylic end groups, at 3446 cm−1 due to the stretching of the hydroxyl end groups, at 2970 and 2900 cm−1 due to the C–H stretching, 1730 cm−1 due to the C=O stretching of the polyester, 1580 cm−1 due to the stretching vibration of the C=C bonds of the furan ring. The asymmetrical bending of the –CH3 group of the diol appears at 1470 cm−1. The bands of the region 1000–1300 cm−1 are caused by stretching and bending vibrations of the C–O bond of the furan ring and the ester moieties. PIsF and PCHDMF present similar absorption bands, with the exception of the –CH3 peak of PMePF. PCHDMF exhibits very weak absorption values in the region of 3200–3600 cm−1 that corresponds to the –OH and –COOH end groups, which is in agreement with its higher viscosity value reported in Table 1.

The crystalline structure of the polyesters was studied by WAXD. After carrying out a purification process (dissolution in TFA and precipitation by MeOH) followed by annealing (24 h at 100 °C), to achieve the maximum percent crystallinity, only PCHDMF showed crystalline peaks, confessing its semi-crystalline character. Thus, PIsF did not manage to acquire crystallinity, due to endo and exo hydroxyl groups of isosorbide, which prevent the spatial molecule arrangement. Regarding PMePF, the spherulites formation is rather impossible, due to the methyl group of MPD, which disrupts the molecule symmetry and causes steric hindrance. The WAXD diffraction patterns of the polyesters are presented in Figure 3 and it is observed that the PCHDMF pattern includes some crystalline peaks at the 10–35° range. Specifically, PCHDMF shows peaks at 2θ = 10.03, 16.7, 20.19, 22.39 and 30.81°. It seems that there are some similarities in the patterns for PCHDMF and PCHDMT [19].

Figure 1. (a) Structures; and (b) 1H-NMR spectra of PMePF, PIsF and PCHDMF polyesters.

The isosorbide part of PIsF shows a more complicated spectrum due to its spatial configuration.The most deprotected are the “f” protons owing to their “next to ester bond” location and appearat 5.76 ppm (2 H, q). The endo and exo “g” protons are more deprotected than the exo and endo “h”protons, due to the presence of endo and exo ester bond and appear at 5.76 ppm (2 H, t) and 5.45 ppm(2 H, t), respectively. The “i” protons show a different peak, 4.33 ppm (1 H, d) and 4.50 ppm (1 H, d), asthey are characterized as exo protons and are in varying degree deprotected by the endo and exo esterbonds, respectively.

At the PCHDMF spectrum, the “k” protons correspond to two double peaks at 4.38 (2 H, d) and4.5 ppm (2 H, d), respectively. This phenomenon is attributed to the fact that the used CHMD was acis-trans mixture. Thus, the axial “k” protons at trans case interact different than the axial-equatorial“k” protons at cis case. The “l” protons show a multiple peak at 2.00 ppm (2H, m) and the “m” protonsshow a double peak at 1.20 ppm (2 H, d). The absent of any other peaks indicates that polyester withhigh purity have been prepared.

FTIR spectra of the three synthesized polyesters are presented in Figure 2. The spectrum of PMePFshows characteristic peaks at 3550 cm−1 due to the stretching of the O–H stretch of the carboxylic endgroups, at 3446 cm−1 due to the stretching of the hydroxyl end groups, at 2970 and 2900 cm−1 dueto the C–H stretching, 1730 cm−1 due to the C=O stretching of the polyester, 1580 cm−1 due to thestretching vibration of the C=C bonds of the furan ring. The asymmetrical bending of the –CH3 groupof the diol appears at 1470 cm−1. The bands of the region 1000–1300 cm−1 are caused by stretchingand bending vibrations of the C–O bond of the furan ring and the ester moieties. PIsF and PCHDMFpresent similar absorption bands, with the exception of the –CH3 peak of PMePF. PCHDMF exhibitsvery weak absorption values in the region of 3200–3600 cm−1 that corresponds to the –OH and –COOHend groups, which is in agreement with its higher viscosity value reported in Table 1.

The crystalline structure of the polyesters was studied by WAXD. After carrying out a purificationprocess (dissolution in TFA and precipitation by MeOH) followed by annealing (24 h at 100 ◦C), toachieve the maximum percent crystallinity, only PCHDMF showed crystalline peaks, confessing itssemi-crystalline character. Thus, PIsF did not manage to acquire crystallinity, due to endo and exohydroxyl groups of isosorbide, which prevent the spatial molecule arrangement. Regarding PMePF,the spherulites formation is rather impossible, due to the methyl group of MPD, which disrupts themolecule symmetry and causes steric hindrance. The WAXD diffraction patterns of the polyesters arepresented in Figure 3 and it is observed that the PCHDMF pattern includes some crystalline peaksat the 10–35◦ range. Specifically, PCHDMF shows peaks at 2θ = 10.03, 16.7, 20.19, 22.39 and 30.81◦.It seems that there are some similarities in the patterns for PCHDMF and PCHDMT [19].

Materials 2017, 10, 801 6 of 19Materials 2017, 10, 801 6 of 18

4000 3500 3000 2500 2000 1500 1000 500

Ab

sorb

ance

Wavenumber (cm-1)

PMePF

PIsF

PCHDMF

Figure 2. FTIR spectra of the polyesters.

10 20 30 40 50 60

Inte

nsi

ty (

cou

nts

)

2theta (degree)

PMePF

PIsF

PCHDMF

Figure 3. WAXD diffractograms of PMePF, PIsF and PCHDMeF.

3.3. Thermal Characterization

The DSC study (Figure 4) revealed the fact that the PCHDMF is a semi-crystalline material, while PIsF and PMePF are amorphous, as was already found by WAXD (Figure 3). PCHDMF showed a high melting temperature (Tm = 262 °C) measured from the first heating scan (Figure 4a). This value is quite higher than that for poly(ethylene furanoate) (PEF) (Tm = 220 °C), but lower compared to that for poly(cyclohexane dimethylene terephthalate) (PCHDMT) [18]. In the DSC trace of the quenched sample a Tg = 74 °C was found. A sharp cold crystallization peak appeared at 121 °C, showing that PCHDMF is a fast crystallizing polymer. This behavior is quite different to those of the most important and most studied furanoate polyesters, poly(ethylene furanoate) (PEF) and poly(propylene furanoate), or even poly(butylene furanoate) (PBF), which crystallize slowly. Only those furanoates from linear aliphatic diols with even number of methylene groups, exceeding six, show fast crystallization, however, their melting temperatures and Tgs are much lower compared to PCHDMF [20,21]. PIsF shows a Tg = 157 °C, which is lower than that reported in a previous work [22], that was synthesized at lower temperature, using dichlorfurandicarboxylate as monomer in a water/DCM system, achieving higher Mn. However, PIsF is considered an alternative to polycarbonate (PC). This Tg value is comparable and even higher than that of PC (Figure 4b). The thermogram of as received PMePF sample revealed only a glass transition, in contrast to PPF and poly(dimethyl propylene furanoate) (PDMEPF) which are crystallizable and showed well defined melting peaks [23]. For PMePF a Tg = 55 °C was evidenced, which is slightly higher than that of PPF,

Figure 2. FTIR spectra of the polyesters.

Materials 2017, 10, 801 6 of 18

4000 3500 3000 2500 2000 1500 1000 500

Ab

sorb

ance

Wavenumber (cm-1)

PMePF

PIsF

PCHDMF

Figure 2. FTIR spectra of the polyesters.

10 20 30 40 50 60

Inte

nsi

ty (

cou

nts

)

2theta (degree)

PMePF

PIsF

PCHDMF

Figure 3. WAXD diffractograms of PMePF, PIsF and PCHDMeF.

3.3. Thermal Characterization

The DSC study (Figure 4) revealed the fact that the PCHDMF is a semi-crystalline material, while PIsF and PMePF are amorphous, as was already found by WAXD (Figure 3). PCHDMF showed a high melting temperature (Tm = 262 °C) measured from the first heating scan (Figure 4a). This value is quite higher than that for poly(ethylene furanoate) (PEF) (Tm = 220 °C), but lower compared to that for poly(cyclohexane dimethylene terephthalate) (PCHDMT) [18]. In the DSC trace of the quenched sample a Tg = 74 °C was found. A sharp cold crystallization peak appeared at 121 °C, showing that PCHDMF is a fast crystallizing polymer. This behavior is quite different to those of the most important and most studied furanoate polyesters, poly(ethylene furanoate) (PEF) and poly(propylene furanoate), or even poly(butylene furanoate) (PBF), which crystallize slowly. Only those furanoates from linear aliphatic diols with even number of methylene groups, exceeding six, show fast crystallization, however, their melting temperatures and Tgs are much lower compared to PCHDMF [20,21]. PIsF shows a Tg = 157 °C, which is lower than that reported in a previous work [22], that was synthesized at lower temperature, using dichlorfurandicarboxylate as monomer in a water/DCM system, achieving higher Mn. However, PIsF is considered an alternative to polycarbonate (PC). This Tg value is comparable and even higher than that of PC (Figure 4b). The thermogram of as received PMePF sample revealed only a glass transition, in contrast to PPF and poly(dimethyl propylene furanoate) (PDMEPF) which are crystallizable and showed well defined melting peaks [23]. For PMePF a Tg = 55 °C was evidenced, which is slightly higher than that of PPF,

Figure 3. WAXD diffractograms of PMePF, PIsF and PCHDMeF.

3.3. Thermal Characterization

The DSC study (Figure 4) revealed the fact that the PCHDMF is a semi-crystalline material, whilePIsF and PMePF are amorphous, as was already found by WAXD (Figure 3). PCHDMF showed a highmelting temperature (Tm = 262 ◦C) measured from the first heating scan (Figure 4a). This value isquite higher than that for poly(ethylene furanoate) (PEF) (Tm = 220 ◦C), but lower compared to that forpoly(cyclohexane dimethylene terephthalate) (PCHDMT) [18]. In the DSC trace of the quenched samplea Tg = 74 ◦C was found. A sharp cold crystallization peak appeared at 121 ◦C, showing that PCHDMFis a fast crystallizing polymer. This behavior is quite different to those of the most important and moststudied furanoate polyesters, poly(ethylene furanoate) (PEF) and poly(propylene furanoate), or evenpoly(butylene furanoate) (PBF), which crystallize slowly. Only those furanoates from linear aliphaticdiols with even number of methylene groups, exceeding six, show fast crystallization, however,their melting temperatures and Tgs are much lower compared to PCHDMF [20,21]. PIsF shows aTg = 157 ◦C, which is lower than that reported in a previous work [22], that was synthesized at lowertemperature, using dichlorfurandicarboxylate as monomer in a water/DCM system, achieving higherMn. However, PIsF is considered an alternative to polycarbonate (PC). This Tg value is comparableand even higher than that of PC (Figure 4b). The thermogram of as received PMePF sample revealedonly a glass transition, in contrast to PPF and poly(dimethyl propylene furanoate) (PDMEPF) whichare crystallizable and showed well defined melting peaks [23]. For PMePF a Tg = 55 ◦C was evidenced,which is slightly higher than that of PPF, but 13 ◦C lower in comparison to PDMEPF. The addition ofsubstituents methyl groups, gradually decreases the mobility to the macromolecular chains.

Materials 2017, 10, 801 7 of 19

Materials 2017, 10, 801 7 of 18

but 13 °C lower in comparison to PDMEPF. The addition of substituents methyl groups, gradually decreases the mobility to the macromolecular chains.

0 50 100 150 200 250 300

Tcc = 121oC

Tg = 74oC

Tm = 262oC

No

rmal

ize

d H

eat

Flo

w (

W/g

)

Temperature (oC)

PCHDMF (a)

0 50 100 150 200 250 300

Tg=155oC

Hea

t F

low

(m

W)

En

do

Up

Temperature (oC)

PIsF

PC

Tg=157oC

(b)

0 50 100 150 200 250 300

Hea

t F

low

(m

W)

En

do

Up

Temperature (oC)

As received

PDMePF

PMePF

PPF

(c)

-50 0 50 100 150 200 250 300

PDMePF

Hea

t F

low

(m

W)

En

do

Up

Temperature (oC)

Quenched (d)

PPF

PMePF

Figure 4. DSC scans of the prepared furanoate polyesters: (a) for as received and melt quenched PCHDMF; (b) for PIsF and PC; and (c,d) for as received and melt quenched PPF, PMF, and PDMePF.

3.4. Prediction of Dynamic Fragility through Isoconversional and Standard Methods

3.4.1. Variation of the Effective Activation Energy throughout the Glass Transition Temperature

The variation of the effective activation energy (ΔEX) through the glass transition can be determined as a function of the extent of conversion using isoconversional methods. According to the Flynn-Wall-Ozawa (FWO) method [24,25], the value of ΔEX at each value of conversion is calculated as shown in the following equation. = − (ln( ))(1/ ) (1)

where β is the cooling rate, R is the universal gas constant and T the temperature where the conversion X is attained.

ΔEX can be measured upon cooling through the glass transition using the above equation as thermodynamic properties such as Cp exhibit a monotonic decrease upon transformation from an equilibrium liquid to a glass; however, when measured on heating, Cp often shows an overshoot near the glass transition and the size of the enthalpic overshoot is influenced by thermal history and the heating rate [26]. To estimate the effective activation energy through the glass transition region using an isoconversional analysis, the conversion, X, has to be defined. The extent of conversion, X, needed for the computations is determined as the normalized heat capacity, CN

P , that is determined from DSC scans as

Figure 4. DSC scans of the prepared furanoate polyesters: (a) for as received and melt quenchedPCHDMF; (b) for PIsF and PC; and (c,d) for as received and melt quenched PPF, PMF, and PDMePF.

3.4. Prediction of Dynamic Fragility through Isoconversional and Standard Methods

3.4.1. Variation of the Effective Activation Energy throughout the Glass Transition Temperature

The variation of the effective activation energy (∆EX) through the glass transition can bedetermined as a function of the extent of conversion using isoconversional methods. Accordingto the Flynn-Wall-Ozawa (FWO) method [24,25], the value of ∆EX at each value of conversion iscalculated as shown in the following equation.(

∆EXR

)x= −

(d(ln(β))

d(1/T)

)x

(1)

where β is the cooling rate, R is the universal gas constant and T the temperature where the conversionX is attained.

∆EX can be measured upon cooling through the glass transition using the above equation asthermodynamic properties such as Cp exhibit a monotonic decrease upon transformation from anequilibrium liquid to a glass; however, when measured on heating, Cp often shows an overshoot nearthe glass transition and the size of the enthalpic overshoot is influenced by thermal history and theheating rate [26]. To estimate the effective activation energy through the glass transition region usingan isoconversional analysis, the conversion, X, has to be defined. The extent of conversion, X, neededfor the computations is determined as the normalized heat capacity, CN

p , that is determined from DSCscans as

CNp =

(Cp − Cpg

)T(

Cpe − Cpg)

T≡ X (2)

Materials 2017, 10, 801 8 of 19

where Cpg and Cpe refer to the glassy and the equilibrium (liquid) heat capacities, respectively. Tg iscalculated from DSC measurements as the temperature where the step change in CN

p attains half thevalue of the total change.

Note that this definition is the same with that used by Vyazovkin et al. [10,27] while opposite tothat used by Badrianarayanan et al. [26,28] for evaluating cooling curves where X = 0 is taken to be theliquid and X = 1 is taken to be the glass.

Figure 5 displays DSC data taken in the vicinity of Tg at different heating rates. By transformingDSC data to the normalized heat capacity, CN

p vs. T data sets similar to that shown in Figure 6are obtained.

Application of the FWO isoconversional method to the CNp vs. T data yields the effective activation

energy at different degrees of X. Typical plots of ln(β) vs. 1/T at different Xs appear in Figure 7.The estimated values of ∆Ex using two different heating rate regimes are plotted as a function of Xin Figure 8. When employing the full heating range the values increased with X, while when using10–20 ◦C/min the dependency demonstrated a decrease with the extent of conversion from the glassyto the liquid state that is in agreement with literature findings [29].

Materials 2017, 10, 801 8 of 18

= ( − )( − ) ≡ (2)

where Cpg and Cpe refer to the glassy and the equilibrium (liquid) heat capacities, respectively. Tg is calculated from DSC measurements as the temperature where the step change in CN

P attains half the value of the total change.

Note that this definition is the same with that used by Vyazovkin et al. [10,27] while opposite to that used by Badrianarayanan et al. [26,28] for evaluating cooling curves where X = 0 is taken to be the liquid and X = 1 is taken to be the glass.

Figure 5 displays DSC data taken in the vicinity of Tg at different heating rates. By transforming DSC data to the normalized heat capacity, CN

P vs. T data sets similar to that shown in Figure 6 are obtained.

Application of the FWO isoconversional method to the CN P vs. T data yields the effective

activation energy at different degrees of X. Typical plots of ln(β) vs. 1/T at different Xs appear in Figure 7. The estimated values of ΔEx using two different heating rate regimes are plotted as a function of X in Figure 8. When employing the full heating range the values increased with X, while when using 10–20 °C/min the dependency demonstrated a decrease with the extent of conversion from the glassy to the liquid state that is in agreement with literature findings [29].

Figure 5. DSC scans of: PMEPF (a); and PIF (b) glass transition temperature at various heating rates. Figure 5. DSC scans of: PMEPF (a); and PIF (b) glass transition temperature at various heating rates.

Materials 2017, 10, 801 9 of 19Materials 2017, 10, 801 9 of 18

30 40 50 60 70 80 90

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Cp

N

Temperature (°C)

5 oC/min

10 oC/min

15 oC/min

20 oC/min

Figure 6. Normalized heat capacity vs. temperature for PMEPF at various heating rates.

Figure 7. Typical plots for the estimation of the effective activation energy at different relative conversion values.

The ΔEX dependencies can be further converted to the dependencies of ΔEX on T. This is accomplished by replacing X with an average of the temperatures corresponding to this X at different heating rates [10,27]. The resulting temperature dependencies are presented in Figure 9.

0.0 0.2 0.4 0.6 0.8 1.0

200

400

600

800

1000

1200

5-20 oC/min

10-20 oC/min

∆E

X (

kJ/m

ol)

X Figure 8. Variation of the effective activation energy with conversion at two different heating rate regimes (i.e., 5–20 and 10–20 °C/min).

Figure 6. Normalized heat capacity vs. temperature for PMEPF at various heating rates.

Materials 2017, 10, 801 9 of 18

30 40 50 60 70 80 90

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Cp

N

Temperature (°C)

5 oC/min

10 oC/min

15 oC/min

20 oC/min

Figure 6. Normalized heat capacity vs. temperature for PMEPF at various heating rates.

Figure 7. Typical plots for the estimation of the effective activation energy at different relative conversion values.

The ΔEX dependencies can be further converted to the dependencies of ΔEX on T. This is accomplished by replacing X with an average of the temperatures corresponding to this X at different heating rates [10,27]. The resulting temperature dependencies are presented in Figure 9.

0.0 0.2 0.4 0.6 0.8 1.0

200

400

600

800

1000

1200

5-20 oC/min

10-20 oC/min

∆E

X (

kJ/m

ol)

X Figure 8. Variation of the effective activation energy with conversion at two different heating rate regimes (i.e., 5–20 and 10–20 °C/min).

Figure 7. Typical plots for the estimation of the effective activation energy at different relativeconversion values.

The ∆EX dependencies can be further converted to the dependencies of ∆EX on T. This isaccomplished by replacing X with an average of the temperatures corresponding to this X at differentheating rates [10,27]. The resulting temperature dependencies are presented in Figure 9.

Materials 2017, 10, 801 9 of 18

30 40 50 60 70 80 90

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Cp

N

Temperature (°C)

5 oC/min

10 oC/min

15 oC/min

20 oC/min

Figure 6. Normalized heat capacity vs. temperature for PMEPF at various heating rates.

Figure 7. Typical plots for the estimation of the effective activation energy at different relative conversion values.

The ΔEX dependencies can be further converted to the dependencies of ΔEX on T. This is accomplished by replacing X with an average of the temperatures corresponding to this X at different heating rates [10,27]. The resulting temperature dependencies are presented in Figure 9.

0.0 0.2 0.4 0.6 0.8 1.0

200

400

600

800

1000

1200

5-20 oC/min

10-20 oC/min

∆E

X (

kJ/m

ol)

X Figure 8. Variation of the effective activation energy with conversion at two different heating rate regimes (i.e., 5–20 and 10–20 °C/min).

Figure 8. Variation of the effective activation energy with conversion at two different heating rateregimes (i.e., 5–20 and 10–20 ◦C/min).

Materials 2017, 10, 801 10 of 19

Materials 2017, 10, 801 10 of 18

49 50 51 52 53 54 55

200

400

600

800

1000

1200

∆E

X (

kJ/m

ol)

Temperature (oC)

5-20 oC/min

10-20 oC/min

Figure 9. Variation of the effective activation energy with temperature at two different heating rate regimes (i.e., 5–20 and 10–20 °C/min).

Vyazovkin et al. [27], to correlate the effective activation energy with the dynamic fragility of a sample, introduced a variability parameter, ΔE, denoting the rate of change of ΔEX with temperature and defined as = 0.25 − 0.750.25 − 0.75 (3)

where ΔE0.25 and ΔE0.75 are the effective activation energy values at X = 0.25 and 0.75, respectively; and T0.25 and T0.75 are the values of TX for the respective values of X. Then, from the values reported in Figure 9, the variability parameter for both polymers can be estimated, as reported in Table 2.

Table 2. Fragility parameters for different polymers.

Material ΔETg (kJ/mol) Tg (°C) m D ΔE (kJ/mol/K) PMePF 799 ± 14 53.0 128 5.26 −82.6

PIsF 810 ± 46 161.9 97.2 7.26 −32.4 PET [10] 156 −110.6 PEN [29] 451 123 60 13.5 −15.6

Moreover, other than isoconversional methods expect the value of the effective activation energy to remain constant throughput the glass transition. In order to determine thus an average effective activation energy ΔEave, Moynihan et al. [29] have proposed the use of a dependence of the Tg on the rate of heating or cooling in accord with Equation (1) as = ( | |)1/ (4)

For heating, Equation (4) is applicable subject to the constraint that, prior to heating, the glassy material should be cooled from above to well below the glass transition region at a rate whose absolute value is equal to the rate of heating.

In Figure 6, the glass transition temperatures were measured at different heating rates and, using Equation (4), the average effective activation energy was estimated using either the 5 to 20 °C/min heating rate rang or the 10 to 20 °C/min (Figure 10). The resultant values were similar, and are plotted as straight lines in Figures 4 and 5. These follow the corresponding values estimated from the isoconversional methods.

Figure 9. Variation of the effective activation energy with temperature at two different heating rateregimes (i.e., 5–20 and 10–20 ◦C/min).

Vyazovkin et al. [27], to correlate the effective activation energy with the dynamic fragility of asample, introduced a variability parameter, ∆E, denoting the rate of change of ∆EX with temperatureand defined as

∆E =∆E0.25 − ∆E0.75

T0.25 − T0.75(3)

where ∆E0.25 and ∆E0.75 are the effective activation energy values at X = 0.25 and 0.75, respectively;and T0.25 and T0.75 are the values of TX for the respective values of X. Then, from the values reportedin Figure 9, the variability parameter for both polymers can be estimated, as reported in Table 2.

Table 2. Fragility parameters for different polymers.

Material ∆ETg (kJ/mol) Tg (◦C) m D ∆E (kJ/mol/K)

PMePF 799 ± 14 53.0 128 5.26 −82.6PIsF 810 ± 46 161.9 97.2 7.26 −32.4

PET [10] 156 −110.6PEN [29] 451 123 60 13.5 −15.6

Moreover, other than isoconversional methods expect the value of the effective activation energyto remain constant throughput the glass transition. In order to determine thus an average effectiveactivation energy ∆Eave, Moynihan et al. [29] have proposed the use of a dependence of the Tg on therate of heating or cooling in accord with Equation (1) as

∆Eave

R=

d(ln|β|)d(1/Tg

) (4)

For heating, Equation (4) is applicable subject to the constraint that, prior to heating, the glassymaterial should be cooled from above to well below the glass transition region at a rate whose absolutevalue is equal to the rate of heating.

In Figure 6, the glass transition temperatures were measured at different heating rates and, usingEquation (4), the average effective activation energy was estimated using either the 5 to 20 ◦C/minheating rate rang or the 10 to 20 ◦C/min (Figure 10). The resultant values were similar, and areplotted as straight lines in Figures 4 and 5. These follow the corresponding values estimated from theisoconversional methods.

Materials 2017, 10, 801 11 of 19Materials 2017, 10, 801 11 of 18

0.3680 0.3684 0.3688 0.3692 0.3696 0.37001.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

3.0

3.2

ln(β

)

1000/(RTg) (mol/J)

slope = -799 kJ/mol

R2=0.9996

Figure 10. Estimation of the average activation energy at the glass transition for PMePF.

3.4.2. Prediction of the Dynamic Fragility

The dynamic fragility of PMEPF and PIF were estimated using the scanning rate dependency of Tg method, reviewed in the paper of Crowley and Zografi [30].

Accordingly, at a single temperature, the fragility parameter, m, could be defined by [30]:

≡ log = (ln 10) (5)

where τ is a mean relaxation time given by the following form of the Vogel–Tammann–Fulcher (VTF) equation: = exp − (6)

τ0, D and T0 are constants, with D termed the strength parameter, with a large value (>30) representing “strong” behavior and low D value (<10) representing “fragile” behavior. According to the assumptions made in Ref. [30], parameter D can be calculated by Equation (7) using mmin = 16. = (ln 10)− (7)

In Equation (5), ΔETg is equal to the average effective activation energy, ΔEave calculated from the variation of Tg with the heating rate, Equation (4). Therefore, using the value estimated in the previous section for ΔETg, m is calculated from Equation (5) and D from Equation (7). All these values appear in Table 1. A large m value indicates rapidly changing dynamics at Tg which equates to “fragile” behavior. Therefore, it seems that PMEPF is more fragile than PIF and both are much more fragile compared to PEN. However, PMePF seems to have a behavior similar to PET, with the latter being slightly more fragile. The average effective activation energy for both PMePF and PIF are similar though the large different in the average Tg results in a different behavior concerning fragility. According to the D values, all polymers are considered fragile.

3.5. Thermal Degradation

Thermal stability of the polyesters was studied by means of TGA, and the respective thermogravimetric (TG) and differential TG (dTG) curves are presented in Figure 11. As can be seen, all polyesters decompose in two steps, while the maximum rate of decomposition (stage of carbonization) appears at Td.max = 408.9 °C, 421.9 °C and 412.00 °C for PMePF, PIsF and PCHDMF, respectively, while a secondary mass loss takes place at 516.63 °C, 518.49 °C and 526.75 °C respectively, as follows from dTG. It is clear that all polymers are thermally stable materials, which was also proven for other furan polyesters in our previous studies [31,32].

Figure 10. Estimation of the average activation energy at the glass transition for PMePF.

3.4.2. Prediction of the Dynamic Fragility

The dynamic fragility of PMEPF and PIF were estimated using the scanning rate dependency ofTg method, reviewed in the paper of Crowley and Zografi [30].

Accordingly, at a single temperature, the fragility parameter, m, could be defined by [30]:

m ≡ d log τ

d(

TgT

)∣∣∣∣∣∣T=Tg

=∆ETg

(ln 10)RTg(5)

where τ is a mean relaxation time given by the following form of the Vogel–Tammann–Fulcher(VTF) equation:

τ = τ0 exp(

DT0

T − T0

)(6)

τ0, D and T0 are constants, with D termed the strength parameter, with a large value (>30) representing“strong” behavior and low D value (<10) representing “fragile” behavior. According to the assumptionsmade in Ref. [30], parameter D can be calculated by Equation (7) using mmin = 16.

D =(ln 10)m2

minm−mmin

(7)

In Equation (5), ∆ETg is equal to the average effective activation energy, ∆Eave calculated from thevariation of Tg with the heating rate, Equation (4). Therefore, using the value estimated in the previoussection for ∆ETg, m is calculated from Equation (5) and D from Equation (7). All these values appear inTable 1. A large m value indicates rapidly changing dynamics at Tg which equates to “fragile” behavior.Therefore, it seems that PMEPF is more fragile than PIF and both are much more fragile compared toPEN. However, PMePF seems to have a behavior similar to PET, with the latter being slightly morefragile. The average effective activation energy for both PMePF and PIF are similar though the largedifferent in the average Tg results in a different behavior concerning fragility. According to the Dvalues, all polymers are considered fragile.

3.5. Thermal Degradation

Thermal stability of the polyesters was studied by means of TGA, and the respectivethermogravimetric (TG) and differential TG (dTG) curves are presented in Figure 11. As can beseen, all polyesters decompose in two steps, while the maximum rate of decomposition (stage ofcarbonization) appears at Td.max = 408.9 ◦C, 421.9 ◦C and 412.00 ◦C for PMePF, PIsF and PCHDMF,respectively, while a secondary mass loss takes place at 516.63 ◦C, 518.49 ◦C and 526.75 ◦C respectively,

Materials 2017, 10, 801 12 of 19

as follows from dTG. It is clear that all polymers are thermally stable materials, which was also provenfor other furan polyesters in our previous studies [31,32].Materials 2017, 10, 801 12 of 18

100 200 300 400 500 6000

20

40

60

80

100

Mas

s (%

)

Temperature (oC)

(a) PMePF

DG

T

408.9 oC

100 200 300 400 500 6000

20

40

60

80

100

Ma

ss (

%)

Temperature (oC)

(b) PIsF

DT

G

421 oC

100 200 300 400 500 6000

20

40

60

80

100

Mas

s (

%)

Temperature (oC)

(c) PCHDMF

DT

G

412 oC

Figure 11. Remaining mass (TG) and differential (dTG) curves of (a) PMePF, (b) PIsF and (c) PCHDMF.

3.6. Decomposition Mechanism Study by Py/GC-MS

To study the decomposition mechanism of the prepared polyesters in detail, characterization with Py/GC-MS was employed. Initially, EGA was performed and the flash pyrolysis temperature was selected based on the resulting thermograms of Figure 12. The evolution of gases for the polyester

Figure 11. Remaining mass (TG) and differential (dTG) curves of (a) PMePF, (b) PIsF and (c) PCHDMF.

Materials 2017, 10, 801 13 of 19

3.6. Decomposition Mechanism Study by Py/GC-MS

To study the decomposition mechanism of the prepared polyesters in detail, characterization withPy/GC-MS was employed. Initially, EGA was performed and the flash pyrolysis temperature wasselected based on the resulting thermograms of Figure 12. The evolution of gases for the polyesterstarts above 325 ◦C, and ends after 500 ◦C. Therefore, the selected temperature for flash pyrolysis was600 ◦C, where all polyesters were fully decomposed. The resulting chromatographs, after pyrolysisat 600 ◦C are presented in Figure 13. Each peak corresponds to a mass spectrum that correspondsto a decomposition product, and the most important identified compounds for each polyester arepresented in Tables 3–5.

Materials 2017, 10, 801 13 of 18

starts above 325 °C, and ends after 500 °C. Therefore, the selected temperature for flash pyrolysis was 600 °C, where all polyesters were fully decomposed. The resulting chromatographs, after pyrolysis at 600 °C are presented in Figure 13. Each peak corresponds to a mass spectrum that corresponds to a decomposition product, and the most important identified compounds for each polyester are presented in Tables 3–5.

Understanding the decomposition mechanism of polyesters is of great importance since it can help with choosing the most suitable methods or fillers to enhance their thermal stability. Pyrolysis of several furanic polyesters has been widely studied by our research team [23,31,33,34], while the case of PMePF, PIsF and PCHDMF is reported for the first time in this study.

200 250 300 350 400 450 500 550 600 650

Rel

ati

ve

inte

nsi

ty (

%)

Temperature (oC)

PMePF

PIsF

PCHDMF

Figure 12. EGA analysis of the polyesters.

0 5 10 15 20 25 30

Rel

ativ

e In

ten

sity

(%

)

Retention time (min)

PMePF

PIsF

PCHDMF

Figure 13. Gas chromatographs of the polyesters after pyrolysis at 600 °C.

In general, furanic polyesters degrade similarly to terephthalate polyesters, with β-hydrogen heterolytic scission being the dominant mechanism when β-hydrogens are present in the structure of the macromolecular chains. Secondary mechanisms that occur mostly at elevated temperatures are α-scission reactions and homolytic processes.

In small retention times (Rt), volatile molecules of low molecular weights (MW), such as carbon dioxide, furan and 2-furoic acid were identified. As the Rt increases, bigger and more complex structures, such as dimers, were identified. The major products correspond to compounds with

Figure 12. EGA analysis of the polyesters.

Materials 2017, 10, 801 13 of 18

starts above 325 °C, and ends after 500 °C. Therefore, the selected temperature for flash pyrolysis was 600 °C, where all polyesters were fully decomposed. The resulting chromatographs, after pyrolysis at 600 °C are presented in Figure 13. Each peak corresponds to a mass spectrum that corresponds to a decomposition product, and the most important identified compounds for each polyester are presented in Tables 3–5.

Understanding the decomposition mechanism of polyesters is of great importance since it can help with choosing the most suitable methods or fillers to enhance their thermal stability. Pyrolysis of several furanic polyesters has been widely studied by our research team [23,31,33,34], while the case of PMePF, PIsF and PCHDMF is reported for the first time in this study.

200 250 300 350 400 450 500 550 600 650

Rel

ati

ve

inte

nsi

ty (

%)

Temperature (oC)

PMePF

PIsF

PCHDMF

Figure 12. EGA analysis of the polyesters.

0 5 10 15 20 25 30

Rel

ativ

e In

ten

sity

(%

)

Retention time (min)

PMePF

PIsF

PCHDMF

Figure 13. Gas chromatographs of the polyesters after pyrolysis at 600 °C.

In general, furanic polyesters degrade similarly to terephthalate polyesters, with β-hydrogen heterolytic scission being the dominant mechanism when β-hydrogens are present in the structure of the macromolecular chains. Secondary mechanisms that occur mostly at elevated temperatures are α-scission reactions and homolytic processes.

In small retention times (Rt), volatile molecules of low molecular weights (MW), such as carbon dioxide, furan and 2-furoic acid were identified. As the Rt increases, bigger and more complex structures, such as dimers, were identified. The major products correspond to compounds with

Figure 13. Gas chromatographs of the polyesters after pyrolysis at 600 ◦C.

Materials 2017, 10, 801 14 of 19

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure

1.37 44 CO2

1.56 68

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

1.70 70

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

2.01 90

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

3.00 112

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

9.49 168

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

9.82 166

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

12.56 162

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

13.13 184

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

17.59 264

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

19.62 278

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

23.43 322

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

24.64 378

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

24.88 376

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

26.52 389 unidentified

28.01 420

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

29.84 474

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure

1.38 44 CO2

1.57 68

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

4.26 112

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

Materials 2017, 10, 801 15 of 19

Table 4. Cont.

Rt (min) MW (amu) Possible Structure

5.97 126

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

7.34 128

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

11.02 146

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

12.57 162

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

17.88 180

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

18.74 197

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

24.19 222

Materials 2017, 10, 801 15 of 18

23.43 322

24.64 378

24.88 376

26.52 389 unidentified

28.01 420

29.84 474

Table 4. Identified pyrolysis products of PIsF.

Rt (min) MW (amu) Possible Structure1.38 44 CO2

1.57 68

4.26 112

5.97 126

7.34 128

11.02 146

12.57 162

17.88 180

18.74 197

24.19 222

Understanding the decomposition mechanism of polyesters is of great importance since it canhelp with choosing the most suitable methods or fillers to enhance their thermal stability. Pyrolysis ofseveral furanic polyesters has been widely studied by our research team [23,31,33,34], while the caseof PMePF, PIsF and PCHDMF is reported for the first time in this study.

In general, furanic polyesters degrade similarly to terephthalate polyesters, with β-hydrogenheterolytic scission being the dominant mechanism when β-hydrogens are present in the structure ofthe macromolecular chains. Secondary mechanisms that occur mostly at elevated temperatures areα-scission reactions and homolytic processes.

In small retention times (Rt), volatile molecules of low molecular weights (MW), such as carbondioxide, furan and 2-furoic acid were identified. As the Rt increases, bigger and more complexstructures, such as dimers, were identified. The major products correspond to compounds withvinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions,presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) weredetected. Compounds with methyl-ester end groups resulted from the esterification step of thesynthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of doublepeaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture ofthe two different stereoisomers.

Materials 2017, 10, 801 16 of 19

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure

1.37 44 CO2

1.56 68

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

3.48 108

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

4.01 108

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

5.97 127

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

7.76 144

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

8.11 144

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

15.89 192 unidentified

16.18 112

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

20.16 250

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

20.40 247 unidentified

24.24 296

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

24.40 332

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

24.67 332

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

26.26 372

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

26.43 316 unidentified

27.90 390

Materials 2017, 10, 801 16 of 18

Table 5. Identified pyrolysis products of PCHDMF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

3.48 108

4.01 108

5.97 127

7.76 144

8.11 144

15.89 192 unidentified

16.18 112

20.16 250

20.40 247 unidentified

24.24 296

24.40 332

24.67 332

26.26 372

26.43 316 unidentified

27.90 390

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) and poly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensation method and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF and PMEPF polyesters are amorphous and showed Tg at 157 °C and 55 °C, respectively. PCHDMF is semi-crystalline material with Tm = 262 °C and Tg = 74 °C. All three polyesters exhibit high thermal stability, since they degrade at temperatures higher than 300 °C. The degradation mechanism was explored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. and N.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kinetic fragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

Materials 2017, 10, 801 14 of 18

vinyl- and carboxyl- end groups. Those groups are characteristic products of β-scission reactions, presented in Scheme 1. Additionally, the used monomers (DMFD or MFD, DMP, Is, CHDM) were detected. Compounds with methyl-ester end groups resulted from the esterification step of the synthesis procedure. Regarding the chromatograph pattern of PCHDM, the appearance of double peaks is noteworthy. Those correspond to cis-trans isomers, since the used diol CHDM is a mixture of the two different stereoisomers.

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Table 3. Identified pyrolysis products of PMePF.

Rt (min) MW (amu) Possible Structure1.37 44 CO2

1.56 68

1.70 70

2.01 90

3.00 112

9.49 168

9.82 166

12.56 162

13.13 184

17.59 264

19.62 278

Scheme 1. Heterolytic β-hydrogen scission degradation mechanism.

Materials 2017, 10, 801 17 of 19

4. Conclusions

Poly(isosorbide furanoate) (PIs), poly(methyl-propylene furanoate) (PMePF) andpoly(cyclohanedimethylene furanoate) (PCHDMF) were synthesized using the melt polycondensationmethod and a variation of it, as verified by 1H-NMR spectroscopy and IV measurements. PIsF andPMEPF polyesters are amorphous and showed Tg at 157 ◦C and 55 ◦C, respectively. PCHDMF issemi-crystalline material with Tm = 262 ◦C and Tg = 74 ◦C. All three polyesters exhibit high thermalstability, since they degrade at temperatures higher than 300 ◦C. The degradation mechanism wasexplored. Kinetic fragility was calculated for the amorphous materials.

Author Contributions: G.Z.P. and D.N.B. conceived of and designed the experiments. Z.T., V.T., N.K. andN.D. performed the experiments. Z.T., G.Z.P. and D.S.A. wrote the paper. D.S.A. performed the kineticfragility calculations.

Conflicts of Interest: The authors declare no conflict of interest.

References

1. Liu, B.; Zhang, Z.; Lv, K.; Deng, K.; Duan, H. Efficient aerobic oxidation of biomass-derived5-hydroxymethylfurfural to 2,5-diformylfuran catalyzed by magnetic nanoparticle supported manganeseoxide. Appl. Catal. A Gen. 2014, 472, 64–71. [CrossRef]

2. Zhang, Z.; Liu, B.; Lv, K.; Sun, J.; Deng, K. Aerobic oxidation of biomass derived 5-hydroxymethylfurfuralinto 5-hydroxymethyl-2-furancarboxylic acid catalyzed by a montmorillonite K-10 clay immobilizedmolybdenum acetylacetonate complex. Green Chem. 2014, 16, 2762–2770. [CrossRef]

3. Casanova, O.; Iborra, S.; Corma, A. Biomass into Chemicals: Aerobic Oxidation of5-Hydroxymethyl-2-furfural into 2,5-Furandicarboxylic Acid with Gold Nanoparticle Catalysts.ChemSusChem 2009, 2, 1138–1144. [CrossRef] [PubMed]

4. Burgess, S.K.; Leisen, J.E.; Kraftschik, B.E.; Mubarak, C.R.; Kriegel, R.M.; Koros, W.J. Chain mobility,thermal, and mechanical properties of poly(ethylene furanoate) compared to poly(ethylene terephthalate).Macromolecules 2014, 47, 1383–1391. [CrossRef]

5. Spagnuolo, M.L.; Marini, F.; Sarabia, L.A.; Ortiz, M.C. Migration test of Bisphenol A from polycarbonatecups using excitation-emission fluorescence data with parallel factor analysis. Talanta 2017, 167, 367–378.[CrossRef] [PubMed]

6. Sousa, A.F.; Vilela, C.; Fonseca, A.C.; Matos, M.; Freire, C.S.R.; Gruter, G.-J.M.; Coelho, J.F.J.; Silvestre, A.J.D.Biobased polyesters and other polymers from 2,5-furandicarboxylic acid: A tribute to furan excellency.Polym. Chem. 2015, 6, 5961–5983. [CrossRef]

7. Zhang, D.; Dumont, M.-J. Advances in polymer precursors and bio-based polymers synthesized from5-hydroxymethylfurfural. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 1478–1492. [CrossRef]

8. Papageorgiou, G.Z.; Papageorgiou, D.G.; Terzopoulou, Z.; Bikiaris, D.N. Production of bio-based 2,5-furandicarboxylate polyesters: Recent progress and critical aspects in their synthesis and thermal properties.Eur. Polym. J. 2016, 83, 202–229. [CrossRef]

9. Jiang, M.; Liu, Q.; Zhang, Q.; Ye, C.; Zhou, G. A series of furan-aromatic polyesters synthesized via directesterification method based on renewable resources. J. Polym. Sci. Part A Polym. Chem. 2012, 50, 1026–1036.[CrossRef]

10. Vyazovkin, S.; Sbirrazzuoli, N.; Dranca, I. Variation of the Effective Activation Energy Throughout the GlassTransition. Macromol. Rapid Commun. 2004, 25, 1708–1713. [CrossRef]

11. Ahn, J.S.; Kang, S.M.; Kim, M.K.; Kim, Y.J.; Yoon, K.C.; Park, O.O. Effect of alicyclic monomers on thermalproperties of transparent biodegradable polyesters. Macromol. Res. 2016, 24, 609–616. [CrossRef]

12. Thomas, J.M.; Raja, R. Review: The Materials Chemistry of Inorganic Catalysts. Aust. J. Chem. 2002, 54,551–560. [CrossRef]

13. Liu, Y.; Turner, S.R. Synthesis and properties of cyclic diester based aliphatic copolyesters. J. Polym. Sci. PartA Polym. Chem. 2010, 48, 2162–2169. [CrossRef]

Materials 2017, 10, 801 18 of 19

14. Devaney, R. The structure and electrical properties of poly(1,4-cyclohexylenedimethylene terephthalate),a new linear thermoplastic polyester. In Proceedings of the Annual Report 1962 Conference on ElectricalInsulation, Hershey, PA, USA, 15–17 Octorber 1962.

15. Wang, J.; Liu, X.; Zhang, Y.; Liu, F.; Zhu, J. Modification of poly(ethylene 2,5-furandicarboxylate) with1,4-cyclohexanedimethylene: Influence of composition on mechanical and barrier properties. Polymer (UK)2016, 103, 1–8. [CrossRef]

16. Ullmann, F. Ullmann’s Encyclopedia of Industrial Chemistry, 2nd ed.; Wiley-VCH Verlag GmbH & Co.:Weinheim, Germany, 2003.

17. Tsanaktsis, V.; Papageorgiou, G.Z.; Bikiaris, D.N. A facile method to synthesize high-molecular-weightbiobased polyesters from 2,5-furandicarboxylic acid and long-chain diols. J. Polym. Sci. Part A Polym. Chem.2015, 53, 2617–2632. [CrossRef]

18. Papageorgiou, G.Z.; Tsanaktsis, V.; Bikiaris, D.N. Synthesis of poly(ethylene furandicarboxylate) polyesterusing monomers derived from renewable resources: Thermal behavior comparison with PET and PEN.Phys. Chem. Chem. Phys. 2014, 16, 7946–7958. [CrossRef] [PubMed]

19. Chen, T.; Jiang, G.; Zhang, J. Isothermal crystallization behavior and crystal structure of poly(ethyleneterephthalate-co-1,4-cyclohexylene dimethylene terephthalate) (P(ET/CT)) copolyesters. Cryst. Res. Technol.2014, 49, 232–243. [CrossRef]

20. Papageorgiou, D.G.; Guigo, N.; Tsanaktsis, V.; Exarhopoulos, S.; Bikiaris, D.N.; Sbirrazzuoli, N.;Papageorgiou, G.Z. Fast Crystallization and Melting Behavior of a Long-Spaced Aliphatic FurandicarboxylateBiobased Polyester, Poly(dodecylene 2,5-furanoate). Ind. Eng. Chem. Res. 2016, 55, 5315–5326. [CrossRef]

21. Tsanaktsis, V.; Terzopoulou, Z.; Nerantzaki, M.; Papageorgiou, G.Z.; Bikiaris, D.N. New poly(pentylenefuranoate) and poly(heptylene furanoate) sustainable polyesters from diols with odd methylene groups.Mater. Lett. 2016, 178, 64–67. [CrossRef]

22. Gomes, M.; Gandini, A.; Silvestre, A.J.D.; Reis, B. Synthesis and characterization of poly(2,5-furandicarboxylate)s based on a variety of diols. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 3759–3768.[CrossRef]

23. Tsanaktsis, V.; Terzopoulou, Z.; Exarhopoulos, S.; Bikiaris, D.N.; Achilias, D.S.; Papageorgiou, D.G.;Papageorgiou, G.Z. Sustainable, eco-friendly polyesters synthesized from renewable resources: Preparationand thermal characteristics of poly(dimethyl-propylene furanoate). Polym. Chem. 2015, 6, 8284–8296.[CrossRef]

24. Flynn, J.H.; Wall, L.A. General Treatment of the Thermogravimetry of Polymers. J. Res. Natl. Bur. Stand. APhys. Chem. 1966, 70A, 487–523. [CrossRef]

25. Ozawa, T. A New Method of Analyzing Thermogravimetric Data. Bull. Chem. Soc. Jpn. 1965, 38, 1881–1886.[CrossRef]

26. Badrinarayanan, P.; Zheng, W.; Simon, S.L. Isoconversion analysis of the glass transition. Thermochim. Acta2008, 468, 87–93. [CrossRef]

27. Vyazovkin, S.; Sbirrazzuoli, N.; Dranca, I. Variation in Activation Energy of the Glass Transition for Polymersof Different Dynamic Fragility. Macromol. Chem. Phys. 2006, 207, 1126–1130. [CrossRef]

28. Badrinarayanan, P.; Simon, S.L.; Lyng, R.J.; O’Reilly, J.M. Effect of structure on enthalpy relaxation ofpolycarbonate: Experiments and modeling. Polymer (Guildf) 2008, 49, 3554–3560. [CrossRef]

29. Papageorgiou, G.Z.; Achilias, D.S.; Karayannidis, G.P. Estimation of thermal transitions in poly(ethylenenaphthalate): Experiments and modeling using isoconversional methods. Polymer (Guildf) 2010, 51,2565–2575. [CrossRef]

30. Crowley, K.J.; Zografi, G. The use of thermal methods for predicting glass-former fragility. Thermochim. Acta2001, 380, 79–93. [CrossRef]

31. Terzopoulou, Z.; Tsanaktsis, V.; Nerantzaki, M.; Achilias, D.S.; Vaimakis, T.; Papageorgiou, G.Z.; Bikiaris, D.N.Thermal degradation of biobased polyesters: Kinetics and decomposition mechanism of polyesters from2,5-furandicarboxylic acid and long-chain aliphatic diols. J. Anal. Appl. Pyrolysis 2016, 117, 162–175.[CrossRef]

32. Tsanaktsis, V.; Vouvoudi, E.; Papageorgiou, G.Z.; Papageorgiou, D.G.; Chrissafis, K.; Bikiaris, D.N. Thermaldegradation kinetics and decomposition mechanism of polyesters based on 2,5-furandicarboxylic acid andlow molecular weight aliphatic diols. J. Anal. Appl. Pyrolysis 2015, 112, 369–378. [CrossRef]

Materials 2017, 10, 801 19 of 19

33. Terzopoulou, Z.; Karakatsianopoulou, E.; Kasmi, N.; Majdoub, M.; Papageorgiou, G.Z.; Bikiaris, D.N.Effect of catalyst type on recyclability and decomposition mechanism of poly(ethylene furanoate) biobasedpolyester. J. Anal. Appl. Pyrolysis 2017, in press. [CrossRef]

34. Terzopoulou, Z.; Tsanaktsis, V.; Nerantzaki, M.; Papageorgiou, G.Z.; Bikiaris, D.N. Decompositionmechanism of polyesters based on 2,5-furandicarboxylic acid and aliphatic diols with medium and longchain methylene groups. Polym. Degrad. Stab. 2016, 132, 127–136. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open accessarticle distributed under the terms and conditions of the Creative Commons Attribution(CC BY) license (http://creativecommons.org/licenses/by/4.0/).